+ All Categories
Home > Documents > External boundary control of the motion of a rigid body ...fsueur/GKS.pdf · External boundary...

External boundary control of the motion of a rigid body ...fsueur/GKS.pdf · External boundary...

Date post: 18-Mar-2018
Category:
Upload: leque
View: 214 times
Download: 1 times
Share this document with a friend
37
External boundary control of the motion of a rigid body immersed in a perfect two-dimensional fluid Olivier Glass * ,J´ozsefJ.Kolumb´an , Franck Sueur Abstract We consider the motion of a rigid body due to the pressure of a surrounded two-dimensional irrotational perfect incompressible fluid, the whole system being confined in a bounded domain with an impermeable condition on a part of the external boundary. Thanks to an impulsive control strategy we prove that there exists an appropriate boundary condition on the remaining part of the external boundary (allowing some fluid going in and out the domain) such that the im- mersed rigid body is driven from some given initial position and velocity to some final position and velocity in a given positive time, without touching the exter- nal boundary. The controlled part of the external boundary is assumed to have a nonvoid interior and the final position is assumed to be in the same connected component of the set of possible positions as the initial position. Keywords— Fluid-solid interaction; impulsive control; geodesics; coupled ODE- PDE system; fluid mechanics; Euler equation; control problem; external boundary control. MSC: 76B75, 93C15, 93C20. * CEREMADE, UMR CNRS 7534, Universit´ e Paris-Dauphine, PSL Research University, Place du Mar´ echal de Lattre de Tassigny, 75775 Paris Cedex 16, France CEREMADE, UMR CNRS 7534, Universit´ e Paris-Dauphine, PSL Research University, Place du Mar´ echal de Lattre de Tassigny, 75775 Paris Cedex 16, France Institut de Math´ ematiques de Bordeaux, UMR CNRS 5251, Universit´ e de Bordeaux, 351 cours de la Lib´ eration, F33405 Talence Cedex, France. 1
Transcript

External boundary control of the motion of a rigid body

immersed in a perfect two-dimensional fluid

Olivier Glass∗, Jozsef J. Kolumban †, Franck Sueur‡

Abstract

We consider the motion of a rigid body due to the pressure of a surrounded

two-dimensional irrotational perfect incompressible fluid, the whole system being

confined in a bounded domain with an impermeable condition on a part of the

external boundary. Thanks to an impulsive control strategy we prove that there

exists an appropriate boundary condition on the remaining part of the external

boundary (allowing some fluid going in and out the domain) such that the im-

mersed rigid body is driven from some given initial position and velocity to some

final position and velocity in a given positive time, without touching the exter-

nal boundary. The controlled part of the external boundary is assumed to have

a nonvoid interior and the final position is assumed to be in the same connected

component of the set of possible positions as the initial position.

Keywords— Fluid-solid interaction; impulsive control; geodesics; coupled ODE-

PDE system; fluid mechanics; Euler equation; control problem; external boundary

control.

MSC: 76B75, 93C15, 93C20.

∗CEREMADE, UMR CNRS 7534, Universite Paris-Dauphine, PSL Research University, Place du

Marechal de Lattre de Tassigny, 75775 Paris Cedex 16, France†CEREMADE, UMR CNRS 7534, Universite Paris-Dauphine, PSL Research University, Place du

Marechal de Lattre de Tassigny, 75775 Paris Cedex 16, France‡Institut de Mathematiques de Bordeaux, UMR CNRS 5251, Universite de Bordeaux, 351 cours de

la Liberation, F33405 Talence Cedex, France.

1

Contents

1 Introduction and main result 3

1.1 The model without control . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 The control problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Reformulation of the solid’s equation into an ODE 9

2.1 A reminder of the uncontrolled case . . . . . . . . . . . . . . . . . . . . 9

2.2 Extension to the controlled case . . . . . . . . . . . . . . . . . . . . . . . 13

3 Reduction to the case where the displacement, the velocities and the

circulation are small 15

4 Reduction to an approximate controllability result 17

5 Proof of the approximate controllability result Theorem 5 18

5.1 First step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

5.2 Illustration of the method on a toy model . . . . . . . . . . . . . . . . . 19

5.3 Back to the complete model . . . . . . . . . . . . . . . . . . . . . . . . . 21

5.4 About Remark 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

6 Closeness of the controlled system to the geodesic. Proof of Proposi-

tion 3 23

6.1 Proof of Proposition 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

6.2 Proof of Proposition 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

6.3 Proof of Proposition 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

6.4 Proof of Proposition 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

7 Design of the control according to the solid position. Proof of Propo-

sition 2 28

7.1 The case of a homogeneous disk . . . . . . . . . . . . . . . . . . . . . . . 28

7.2 The case when S0 is not a disk . . . . . . . . . . . . . . . . . . . . . . . 32

2

1 Introduction and main result

1.1 The model without control

A simple model of fluid-solid evolution is that of a single rigid body surrounded by a per-

fect incompressible fluid. Let us describe this system. We consider a two-dimensional

bounded, open, smooth and simply connected1 domain Ω ⊂ R2. The domain Ω is

composed of two disjoint parts: the open part F(t) filled with fluid and the closed part

S(t) representing the solid. These parts depend on time t. Furthermore, we assume

that S(t) is also smooth and simply connected. On the fluid part F(t), the velocity

field u : [0, T ] × F(t) → R2 and the pressure field π : [0, T ] × F(t) → R satisfy the

incompressible Euler equation:

∂u

∂t+ (u · ∇)u+∇π = 0 and div u = 0 for t ∈ [0, T ] and x ∈ F(t). (1.1)

We consider impermeability boundary conditions, namely, on the solid boundary, the

normal velocity coincides with the solid normal velocity

u · n = uS · n on ∂S(t), (1.2)

where uS denotes the solid velocity described below, while on the outer part of the

boundary we have

u · n = 0 on ∂Ω, (1.3)

where n is the unit outward normal vector on ∂F(t). The solid S(t) is obtained by

a rigid movement from S(0), and one can describe its position by the center of mass,

h(t), and the angle variable with respect to the initial position, ϑ(t). Consequently, we

have

S(t) = h(t) +R(ϑ(t))(S0 − h0), (1.4)

where h0 is the center of mass at initial time, and

R(ϑ) =

(cosϑ − sinϑ

sinϑ cosϑ

).

Moreover the solid velocity is hence given by

uS(t, x) = h′(t) + ϑ′(t)(x− h(t))⊥, (1.5)

where for x = (x1, x2) we denote x⊥ = (−x2, x1).

1The condition of simple connectedness is actually not essential and one could generalize the present

result to the case where Ω is merely open and connected at the price of long but straightforward

modifications.

3

The solid evolves according to Newton’s law, and is influenced by the fluid’s pressure

on the boundary:

mh′′(t) =

∫∂S(t)

π n dσ and J ϑ′′(t) =

∫∂S(t)

π (x− h(t))⊥ · ndσ. (1.6)

Here the constants m > 0 and J > 0 denote respectively the mass and the moment

of inertia of the body, where the fluid is supposed to be homogeneous of density 1,

without loss of generality. Furthermore, the circulation around the body is constant in

time, that is ∫∂S(t)

u(t) · τ dσ =

∫∂S0

u0 · τ dσ = γ ∈ R, ∀t ≥ 0, (1.7)

due to Kelvin’s theorem, where τ denotes the unit counterclockwise tangent vector.

τ

∂ΩF(t)

S0

ϑ(t)

τ

n

nh0

S(t)

h(t)

Figure 1: The domains Ω, S(t) and F(t) = Ω \ S(t)

The Cauchy problem for this system with initial data

u|t=0 = u0 for x ∈ F(0),

h(0) = h0, h′(0) = h′0, ϑ(0) = 0, ϑ′(0) = ϑ′0,

(1.8)

is now well-understood, see e.g. [20, 24, 29, 35, 36]. Furthermore, the 3D case has also

been studied in [25, 37]. Note in passing that it is our convention used throughout the

paper that ϑ(0) = 0.

In this paper, we will furthermore assume that the fluid is irrotational at the initial

time, that is curl u0 = 0 in F(0), which implies that it stays irrotational at all times,

due to Helmholtz’s third theorem, i.e.

curl u = 0 for x ∈ F(t), ∀t ≥ 0. (1.9)

1.2 The control problem

We are now in position to state our main result.

Our goal is to investigate the possibility of controlling the solid by means of a

boundary control acting on the fluid. Consider Σ a nonempty, open part of the outer

4

boundary ∂Ω. Suppose that one can choose some non-homogeneous boundary condi-

tions on Σ. One natural possibility is due to Yudovich (see [38]), which consists in

prescribing on the one hand the normal velocity on Σ, i.e. choosing some function

g ∈ C∞0 ([0, T ]× Σ) with∫

Σ g = 0 and imposing that

u(t, x) · n(x) = g(t, x) on [0, T ]× Σ, (1.10)

while on the rest of the boundary we have the usual impermeability condition

u · n = 0 on [0, T ]× (∂Ω \ Σ), (1.11)

and on the other hand the vorticity on the set Σ− of points of [0, T ] × Σ where the

velocity field points inside Ω. Note that Σ− is deduced immediately from g.

Since we are interested in the vorticity-free case, we will actually consider here a

null control in vorticity, that is

curlu(t, x) = 0 on Σ− = (t, x) ∈ [0, T ]× Σ such that u(t, x) · n(x) < 0. (1.12)

Condition (1.12) enforces the validity of (1.9) as in the uncontrolled setting despite the

fact that some fluid is entering the domain.

The general question of this paper is how to control the solid’s movement by using

the above boundary control (that is, the function g). In particular we raise the question

of driving the solid from a given position and a given velocity to some other prescribed

position and velocity. Remark that we cannot expect to control the fluid velocity in

the situation described above: for instance, Kelvin’s theorem gives an invariant of the

dynamics, regardless of the control.

Throughout this paper we will only consider solid trajectories which stay away from

the boundary. Therefore we introduce

Q = q := (h, ϑ) ∈ Ω× R : d(h+R(ϑ)(S0 − h0), ∂Ω) > 0.

The main result of this paper is the following statement.

Theorem 1. Let T > 0. Consider S0 ⊂ Ω bounded, closed, simply connected with

smooth boundary, which is not a disk, and u0 ∈ C∞(F(0);R2), γ ∈ R, q0 = (h0, 0), q1 =

(h1, ϑ1) ∈ Q, h′0, h′1 ∈ R2, ϑ′0, ϑ

′1 ∈ R, such that (h0, 0) and (h1, ϑ1) belong to the same

connected component of Q and

div u0 = curlu0 = 0 in F(0), u0 · n = 0 on ∂Ω,

u0 · n = (h′0 + ϑ′0(x− h0)⊥) · n on ∂S0,

∫∂S0

u0 · τ dσ = γ.

Then there exists a control g ∈ C∞0 ((0, T )×Σ) and a solution (h, ϑ, u) ∈ C∞([0, T ];Q)×C∞([0, T ];C∞(F(t);R2)) to (1.1), (1.2), (1.6), (1.7), (1.8), (1.9), (1.10), (1.11), which

satisfies (h, h′, ϑ, ϑ′)(T ) = (h1, h′1, ϑ1, ϑ

′1).

5

Note that there is a slight abuse of notation in writing C∞([0, T ];C∞(F(t);R2)),

since the domain in which the fluid evolves is also time-dependent.

(h0, 0) (h′1, ϑ′1)

Σ ⊂ ∂Ω

(h′0, ϑ′0)

(h1, ϑ1)

Figure 2: The initial and final positions and velocities in the control problem

Remark 1. In Theorem 1 the control g can be chosen with an arbitrary small total

flux through Σ−, that is for any T > 0, for any ν > 0, there exists a control g and a

solution (h, ϑ, u) satisfying the properties of Theorem 1 and such that moreover∣∣∣∣∫ T

0

∫Σ−

u · n dσdt∣∣∣∣ < ν.

See Section 5.4 for more explanations. Let us mention that such a small flux condition

cannot be guaranteed in the results [7, 14, 16] regarding the controllability of the Euler

equations.

When S0 is a disk, the second equation in (1.6) becomes degenerate, so it needs to

be treated separately. For instance, in the case of a homogeneous disk, i.e. when the

center of mass coincides with the center of the disk and we have (x−h(t))⊥ ·n = 0, for

any x ∈ ∂S(t), t ≥ 0, hence we cannot control ϑ. However, we have a similar result for

controlling the center of mass h.

Theorem 2. Let T > 0. Given a homogeneous disk S0 ⊂ Ω, u0 ∈ C∞(F(0);R2),

γ ∈ R, h0, h1 ∈ Ω, h′0, h′1 ∈ R2, such that (h0, 0) and (h1, 0) are in the same connected

component of Q, and div u0 = curlu0 = 0 in F(0), u0 · n = 0 on ∂Ω, u0 · n = h′0 · non ∂S0,

∫∂S0 u0 · τ dσ = γ, there exists g ∈ C∞0 ((0, T ) × Σ) and a solution (h, u) in

C∞([0, T ]; Ω) × C∞([0, T ];C∞(F(t);R2)) of (1.1), (1.2), (1.6), (1.7), (1.9), (1.10),

(1.11), (1.12) with initial data (h0, h′0, u0), which satisfies (h, h′)(T ) = (h1, h

′1).

The proof is similar to that of Theorem 1, with the added consideration that (x−h(t))⊥ · n = 0, for any x ∈ ∂S(t), t ≥ 0. We therefore omit the proof. In the case

where the disk is non-homogeneous the analysis is technically more intricate already in

the uncontrolled setting, see [21], and we will let aside this case in this paper.

6

References. Let us mention a few results of boundary controllability of a fluid alone,

that is without any moving body. The problem is then finding a boundary control which

steers the fluid velocity from u0 to some prescribed state u1. For the incompressible

Euler equations small-time global exact boundary controllability has been obtained in

[7, 16] in the 2D, respectively 3D case. This result has been recently extended to the case

of the incompressible Navier-Stokes equation with Navier slip-with-friction boundary

conditions in [10], see also [11] for a gentle exposition. Note that the proof there relies

on the previous results for the Euler equations by means of a rapid and strong control

which drives the system in a high Reynolds regime. This strategy was initiated in [8],

where an interior controllability result was already established. For “viscous fluid +

rigid body” control systems (with Dirichlet boundary conditions), local controllability

results have already been obtained in both 2D and 3D, see e.g. [2, 3, 30]. These

results rely on Carleman estimates on the linearized equation, and consequently on the

parabolic character of the fluid equation.

A different type of fluid-solid control result can be found in [22], where the fluid is

governed by the two-dimensional Euler equation. However in this paper the control is

located on the solid’s boundary which makes the situation quite different.

Actually, the results of Theorem 1 and Theorem 2 can rather be seen as some

extensions to the case of an immersed body of the results [17, 18, 19] concerning La-

grangian controllability of the incompressible Euler and Stokes equations, where the

control takes the same form as here.

Generalizations and open problems. First, as we mentioned before, using the

techniques of this paper, the result could be straightforwardly generalized for non simply

connected domains. One could also manage in the same way the control of several solids

(the reader may in particular see that the argument using Runge’s theorem in Section 7

is local around the solid). We would also like to underline that the absence of vorticity

is not central here. This may surprise the reader acquainted with the Euler equation,

but actually following the arguments of Coron [7, 8], one knows how to control the full

model when one can control the irrotational one. This is by the way the technique

that we use to take care of the circulation γ (see in particular Section 3). But the

presence of vorticity makes a lot of complications from the point of view of the initial

boundary problem, in particular for what concerns the uniqueness issue, see Yudovich

[38]. To avoid these unnecessary technical complications, we restrain ourselves to the

irrotational problem. But the full problem could undoubtedly be treated in the same

way.

There remain also many open problems. Considering the recent progresses on the

7

controllability in the viscous case, a natural question is whether or not the results in this

paper could be adapted to the case where a rigid body is moving in a fluid driven by the

incompressible Navier-Stokes equation, with Navier slip-with-friction boundary condi-

tions. We hope that the analysis performed here in the case of “inviscid fluid + rigid

body” control systems could be used in order to get small-time global controllability

results of “viscous fluid + rigid body” control systems.

Let us mention the following open problem regarding the motion planning of a rigid

body immersed in an inviscid incompressible irrotational flow.

Open problem 1. Let T > 0, (h0, 0) in Q, ξ in C2([0, T ];Q), with ξ(0) = (h0, 0).

Let us decompose ξ′(0) into ξ′(0) = (h′0, ϑ′0). Consider S0 ⊂ Ω bounded, closed, simply

connected with smooth boundary, which is not a disk, γ ∈ R, and u0 ∈ C∞(F(0);R2)

such that div u0 = curlu0 = 0 in F(0), u0 ·n = 0 on ∂Ω, u0 ·n = (h′0 +ϑ′0(x−h0)⊥) ·non ∂S0 and

∫∂S0 u0 · τ dσ = γ. Do there exist g ∈ C0([0, T ] × Σ) and a solution

(h, ϑ, u) ∈ C2([0, T ];Q) × C∞([0, T ];C1(F(t);R2)) to (1.1), (1.2), (1.6), (1.7), (1.8),

(1.9), (1.10), (1.11), which satisfies ξ = (h, ϑ)?

Even the approximate motion planning in C2, i.e. the same statement as above but

with ‖ξ − (h, ϑ)‖C2([0,T ]) ≤ ε (with ε > 0 arbitrary) instead of ξ = (h, ϑ), is an open

problem.

Plan of the paper. The paper is organized as follows. In Section 2 we first recall

from [21] a reformulation of the Newton equations (1.6) as an ODE in the uncontrolled

case and then extend it to the case with control. In particular in the case with zero

circulation and no control this ODE is the geodesic equation associated with a metric

which encodes the added mass phenomenon.

In Section 3 we prove that Theorem 1 can be deduced from a simpler result, namely

Theorem 4, where the solid displacement, the initial and final solid velocities and the

circulation are assumed to be small.

In Section 4 we prove that another reduction is possible, as we prove that an ap-

proximate controllability result (rather than an exact one), namely Theorem 5, allows

to deduce Theorem 4.

Section 5 is devoted to the proof of Theorem 5 and is the core of the paper.

In Section 6 we prove a Proposition that is important for Theorem 5, namely that

we can approximate the whole system by a simpler one in a certain regime.

Section 7 explains how one can construct the control by means of complex analysis:

it can be considered as the cornerstone of our control strategy.

8

2 Reformulation of the solid’s equation into an ODE

In this section we establish a reformulation of the Newton equations (1.6) as an ODE

for the three degrees of freedom of the rigid body with coefficients obtained by solving

some elliptic-type problems on a domain depending on the solid position. Indeed the

fluid velocity can be recovered from the solid position and velocity by an elliptic-type

problem, so that the fluid state may be seen as solving an auxiliary steady problem,

where time only appears as a parameter, instead of the evolution equation (1.1). The

Newton equations can therefore be rephrased as a second-order differential equation on

the solid position whose coefficients are determined by the auxiliary fluid problem.

Such a reformulation in the case without boundary control was already achieved in

[21] and we will start by recalling this case in Section 2.1, cf. Proposition 1 below. A

crucial fact in the analysis is that in the ODE reformulation the pre-factor of the body’s

accelerations is the sum of the inertia of the solid and of the so-called “added inertia”

which is a symmetric positive-semidefinite matrix depending only on the body’s shape

and position, and which encodes the amount of incompressible fluid that the rigid body

has also to accelerate around itself. Remarkably enough in the case without control

and where the circulation is 0 it turns out that the solid equations can be recast as a

geodesic equation associated with the metric given by the total inertia.

Then we will extend this analysis to the case where there is a control on a part of the

external boundary in Section 2.2, cf. Theorem 3. In particular we will establish that

the remote influence of the external boundary control translates into two additional

force terms in the second-order ODE for the solid position; indeed we will distinguish

one force term associated with the control velocity and another one associated with its

time derivative.

To simplify notations, we denote the positions and velocities q = (h, ϑ), q′ = (h′, ϑ′),

and

S(q) = h+R(ϑ)(S0 − h0) and F(q) = Ω \ S(q),

since the dependence in time of the domain occupied by the solid comes only from the

position q. Furthermore, we denote q(t) = (h(t), ϑ(t)).

2.1 A reminder of the uncontrolled case

We first recall that in the case without any control the fluid velocity satisfies (1.2),

(1.3), (1.7) and (1.9). Therefore at each time t the fluid velocity u satisfies the following

9

div/curl system:div u = curlu = 0 in F(q),

u · n = 0 on ∂Ω and u · n =(h′ + ϑ′(x− h)⊥

)· n on ∂S(q),∫

∂S(q)u · τ dσ = γ,

(2.1)

where the dependence in time is only due to the one of q and q′. Given the solid position

q and the right hand sides, the system (2.1) uniquely determines the fluid velocity u in

the space of C∞ vector fields on the closure of F(q). Moreover thanks to the linearity

of the system with respect to its right hand sides, its unique solution u can be uniquely

decomposed with respect to the following functions which depend only on the solid

position q = (h, ϑ) in Q and encode the contributions of elementary right hand sides.

• The Kirchhoff potentials

Φ = (Φ1,Φ2,Φ3)(q, ·) (2.2)

are defined as the solution of the Neumann problems

∆Φi(q, x) = 0 in F(q), ∂nΦi(q, x) = 0 on ∂Ω, for i ∈ 1, 2, 3,

∂nΦi(q, x) =

ni on ∂S(q), for i ∈ 1, 2,(x− h)⊥ · n on ∂S(q), for i = 3,

(2.3)

where all differential operators are with respect to the variable x.

• The stream function ψ for the circulation term is defined in the following way.

First we consider the solution ψ(q, ·) of the Dirichlet problem ∆ψ(q, x) = 0 in

F(q), ψ(q, x) = 0 on ∂Ω, ψ(q, x) = 1 on ∂S(q). Then we set

ψ(q, ·) = −

(∫∂S(q)

∂nψ(q, x) dσ

)−1

ψ(q, ·), (2.4)

such that we have ∫∂S(q)

∂nψ(q, x) dσ = −1,

noting that the strong maximum principle gives us ∂nψ(q, x) < 0 on ∂S(q).

Remark 2. The Kirchhoff potentials Φ and the stream function ψ are C∞ as functions

of q on Q. We will use several times some properties of regularity with respect to the

domain of solutions to linear elliptic problems, included for another potential A[q, g]

associated with the control, see Definition 1 below. We will mention along the proof

the properties which will be used and we refer to [6, 28, 32] for more on this material

which is now standard in fluid-structure interaction.

10

The following statement is an immediate consequence of the definitions above.

Lemma 1. For any q = (h, ϑ) in Q, for any p = (`, ω) in R2 × R and for any γ, the

unique solution u in C∞(F(q)) to the following system:div u = curlu = 0 in ∈ F(q),

u · n = 0 on ∂Ω and u · n =(`+ ω(x− h)⊥

)· n on ∂S(q),∫

∂S(q)u · τ dσ = γ.

(2.5)

is given by the following formula, for x in F(q),

u(x) = ∇(p · Φ(q, x)) + γ∇⊥ψ(q, x). (2.6)

Above p · Φ(q, x) denotes the inner product p · Φ(q, x) =∑3

i=1 piΦi(q, x).

Let us now address the solid dynamics. The solid motion is driven by the Newton

equations (1.6) where the influence of the fluid on the solid appears through the fluid

pressure. The pressure can in turn be related to the fluid velocity thanks to the Euler

equations (1.1). The contributions to the solid dynamics of the two terms in the right

hand side of the fluid velocity decomposition formula (2.6) are very different. On the

one hand the potential part, i.e. the first term in the right hand side of (2.6), contributes

as an added inertia matrix, together with a connection term which ensures a geodesic

structure (see [34]), whereas on the other hand the contribution of the term due to the

circulation, i.e. the second term in the right hand side of (2.6), turns out to be a force

which reminds us of the Lorentz force in electromagnetism by its structure (see [21]).

We therefore introduce the following notations.

• We respectively define the genuine and added mass 3× 3 matrices by

Mg =

m 0 0

0 m 0

0 0 J

,

and, for q ∈ Q,

Ma(q) =

(∫F(q)∇Φi(q, x) · ∇Φj(q, x) dx

)16i,j63

.

Note that Ma is a symmetric Gram matrix and is C∞ on Q.

• We define the symmetric bilinear map Γ(q) given by

〈Γ(q), p, p〉 =

∑1≤i,j≤3

Γki,j(q) pi pj

1≤k≤3

∈ R3, ∀p ∈ R3,

11

where, for each i, j, k ∈ 1, 2, 3, Γki,j denotes the Christoffel symbols of the first

kind defined on Q by

Γki,j =1

2

(∂(Ma)k,j

∂qi+∂(Ma)k,i∂qj

− ∂(Ma)i,j∂qk

). (2.7)

It can be checked that Γ is of class C∞ on Q.

• We introduce the following C∞ vector fields on Q with values in R3 by

E = −1

2

∫∂S(q)

|∂nψ(q, ·)|2∂nΦ(q, ·) dσ, (2.8)

B =

∫∂S(q)

∂nψ(q, ·) (∂nΦ(q, ·)× ∂τΦ(q, ·)) dσ. (2.9)

We recall that the notation Φ was given in (2.2).

The reformulation of the model as an ODE is given in the following result, which

was first established in [34] in the case γ = 0 and in [21] in the case γ ∈ R.

Theorem 3. Given q = (h, ϑ) ∈ C∞([0, T ];Q), u ∈ C∞([0, T ];C∞(F(q(t));R2)) we

have that (q, u) is a solution to (1.1), (1.2), (1.3), (1.6), (1.7) and (1.9) if and only if

q satisfies the following ODE on [0, T ](Mg +Ma(q)

)q′′ + 〈Γ(q), q′, q′〉 = γ2E(q) + γq′ ×B(q), (2.10)

and u is the unique solution to the system (2.1). Moreover the total kinetic energy12

(Mg +Ma(q)

)q′ · q′ is conserved in time for smooth solutions of (2.10), at least as

long as there is no collision.

Note that in the case where γ = 0, the ODE (2.10) means that the particle q is

moving along the geodesics associated with the Riemannian metric induced on Q by

the matrix field Mg +Ma(q). Note that, since Q is a manifold with boundary and

the metric Mg +Ma(q) may become singular at the boundary of Q, the Hopf-Rinow

theorem does not apply and geodesics may not be global. However we will make use

only of local geodesics.

Remark 3. Let us also mention that the whole “inviscid fluid + rigid body” system

can be reinterpreted as a geodesic flow on an infinite dimensional manifold, cf. [23].

However the reformulation established by Theorem 3 relies on the finite dimensional

manifold Q and sheds more light on the dynamics of the rigid body.

Below we provide a sketch of the proof of Theorem 3; this will be useful in Section

2.2 when extending the analysis to the controlled case.

12

Proof. Let us focus on the direct part of the proof for sake of clarity but all the subse-

quent arguments can be arranged in order to insure the converse part of the statement

as well. Using Green’s first identity and the properties of the Kirchhoff functions, the

Newton equations (1.6) can be rewritten as

Mg q′′ =

∫F(q)∇π · ∇Φ(q, x) dx. (2.11)

Moreover when u is irrotational, Equation (1.1) can be rephrased as

∇π = −∂tu−1

2∇x|u|2, for x in F(q(t)), (2.12)

and Lemma 1 shows that for any t in [0, T ],

u(t, ·) = ∇(q′(t) · Φ(q(t), ·)) + γ∇⊥ψ(q(t), ·). (2.13)

Substituting (2.13) into (2.12) and then the resulting decomposition of ∇π into (2.11)

we get

Mg q′′ = −

∫F(q)

(∂t∇(q′ · Φ(q, x)) +

∇|∇(q′ · Φ(q, x))|2

2

)· ∇Φ(q, x) dx

−γ∫F(q)

(∂t∇⊥ψ(q, x) +∇

(∇(q′ · Φ(q, x)) · ∇⊥ψ(q, x)

))· ∇Φ(q, x) dx

−γ2

∫F(q)

∇|∇ψ(q, x)|2

2· ∇Φ(q, x) dx.

(2.14)

According to Lemmas 32, 33 and 34 in [21], the terms in the three lines of the right-

hand side above are respectively equal to −Ma(q)q′′ − 〈Γ(q), q′, q′〉, γq′ × B(q) and

γ2E(q), so that we easily deduce the ODE (2.10) from (2.14).

The conservation of the kinetic energy 12

(Mg+Ma(q)

)q′ ·q′ is then simply obtained

by multiplying the ODE (2.10) by q′ and observing that((Mg +Ma(q)

)q′′ + 〈Γ(q), q′, q′〉

)· q′ =

(1

2

(Mg +Ma(q)

)q′ · q′

)′. (2.15)

2.2 Extension to the controlled case

We now tackle the case where a control is imposed on the part Σ of the external

boundary ∂Ω. At any time this control has to be compatible with the incompressibility

of the fluid meaning that the flux through Σ has to be zero. We therefore introduce

the set

C :=

g ∈ C∞0 (Σ;R) such that

∫Σg dσ = 0

.

The decomposition of the fluid velocity u then involves a new potential term in-

volving the following function.

13

Definition 1. With any q ∈ Q and g ∈ C we associate the unique solution α :=

A[q, g] ∈ C∞(F(q);R) to the following Neumann problem:

∆α = 0 in F(q) and ∂n α = g1Σ on ∂F(q), (2.16)

with zero mean on F(q).

Let us mention that the zero mean condition above allows to determine a unique

solution to the Neumann problem but plays no role in the sequel.

Now Lemma 1 can be modified as follows.

Lemma 2. For any q = (h, ϑ) in Q, for any p = (`, ω) in R2 × R, for any g in C, the

unique solution u in C∞(F(q)) to

div u = curlu = 0 in F(q),

u · n = 1Σ g on ∂Ω and u · n =(`+ ω(x− h)⊥

)· n on ∂S(q),∫

∂S(q)u · τ dσ = γ,

is given by

u = ∇(p · Φ(q, ·)) + γ∇⊥ψ(q, ·) +∇A[q, g]. (2.17)

Let us avoid a possible confusion by mentioning that the ∇ operator above has to

be considered with respect to the space variable x. The function A[q, g] and its time

derivative will respectively be involved into the arguments of the following force terms.

Definition 2. We define, for any q in Q, p in R3, α in C∞(F(q);R) and γ in R,

F1(q, p, γ)[α] and F2(q)[α] in R3 by

F1(q, p, γ)[α] := −1

2

∫∂S(q)

|∇α|2 ∂nΦ(q, ·) dσ (2.18)

−∫∂S(q)

∇α ·(∇(p · Φ(q, ·)) + γ∇⊥ψ(q, ·)

)∂nΦ(q, ·) dσ,

F2(q)[α] := −∫∂S(q)

α∂nΦ(q, ·) dσ. (2.19)

Observe that Formulas (2.18) and (2.19) only require α and ∇α to be defined on

∂S(q). Moreover when these formulas are applied to α = A[q, g] for some g in C, then

only the trace of α and the tangential derivative ∂τα on ∂S(q) are involved, since the

normal derivative of α vanishes on ∂S(q) by definition, cf. (2.16).

We define our notion of controlled solution of the “fluid+solid” system as follows.

Definition 3. We say that (q, g) in C∞([0, T ];Q)×C∞0 ([0, T ]; C) is a controlled solution

if the following ODE holds true on [0, T ]:(Mg +Ma(q)

)q′′ + 〈Γ(q), q′, q′〉 = γ2E(q) + γq′ ×B(q)

+ F1(q, q′, γ)[α] + F2(q)[∂tα],(2.20)

where α(t, ·) := A[q(t), g(t, ·)].

14

We have the following result for reformulating the model as an ODE.

Proposition 1. Given

q ∈ C∞([0, T ];Q), u ∈ C∞([0, T ];C∞(F(q(t));R2)) and g ∈ C∞0 ([0, T ]; C),

we have that (q, u) is a solution to (1.1), (1.2), (1.6), (1.7), (1.8), (1.9), (1.10), (1.11),

(1.12) if and only if (q, g) is a controlled solution and u is the unique solution to the

unique div/curl type problem:

div u = curlu = 0 in F(q),

u · n = 1Σ g on ∂Ω and u · n =(h′ + ϑ′(x− h)⊥

)· n on ∂S(q),∫

∂S(q)u · τ dσ = γ,

with q = (h, ϑ).

Proposition 1 therefore extends Theorem 3 to the case with an external bound-

ary control (in particular one recovers Theorem 3 in the case where g is identically

vanishing).

Proof. We proceed as in the proof of Theorem 3 recalled above, with some modifications

due to the extra term involved in the decomposition of the fluid velocity, compare (2.6)

and (2.17). In particular some extra terms appear in the right hand side of (2.14) after

substituting the right hand side of (2.17) for u in (2.12). Using some integration by

parts and the properties of the Kirchhoff functions we obtain integrals on ∂S(q) whose

sum precisely gives F1(q, q′, γ)[α(t, ·)] + F2(q)[∂tα(t, ·)]. This allows to conclude.

3 Reduction to the case where the displacement, the ve-

locities and the circulation are small

For δ > 0, we introduce the set

Qδ = q ∈ Ω× R : d(S(q), ∂Ω) > δ. (3.1)

The goal of this section is to prove that Theorem 1 can be deduced from the following

result. The balls have to be understood for the Euclidean norm (rather than for the

metric Mg +Ma(q)).

Theorem 4. Given δ > 0, S0 ⊂ Ω bounded, closed, simply connected with smooth

boundary, which is not a disk, q0 in Qδ and T > 0, there exists r > 0 such that for

any q1 in B(q0, r), for any γ ∈ R with |γ| ≤ r and for any q′0, q′1 ∈ B(0, r), there is a

controlled solution (q, g) in C∞([0, T ];Qδ)×C∞0 ([0, T ]×Σ) such that (q, q′)(0) = (q0, q′0)

and (q, q′)(T ) = (q1, q′1).

15

Remark in particular that for r > 0 small enough, B(q0, r) is included in the con-

nected component of Qδ containing q0.

Proof of Theorem 1 from Theorem 4. We proceed in two steps: first we use a time-

rescaling argument in order to deduce from Theorem 4 a more general result covering

the case where the initial and final velocities q′0 and q′1 and the circulation γ are large.

This argument is reminiscent of a time-rescaling argument used by J.-M. Coron for

the Euler equation [7], which has been also used in [22] in order to pass from the

potential case to the case with vorticity. Then we use a compactness argument in order

to deal with the case where q0 and q1 are remote (but of course in the same connected

component of Qδ).The time-rescaling argument relies on the following observation: it follows from

(2.20) that (q, g) is a controlled solution on [0, T ] with circulation γ if and only if

(qλ, gλ) is a controlled solution on [0, λT ] with circulation γλ , where (qλ, gλ) is defined

by

qλ(t) := q

(t

λ

)and gλ(t, x) :=

1

λg

(t

λ, x

). (3.2)

Of course the initial and final conditions

(q, q′)(0) = (q0, q′0) and (q, q′)(T ) = (q1, q

′1)

translate respectively into

(qλ, (qλ)′)(0) =

(q0,

q′0λ

)and (qλ, (qλ)′)(T ) =

(q1,

q′1λ

). (3.3)

Now consider q0 in Qδ and q1 in B(q0, r) in the same connected component of Qδas q0, with r > 0 as in Theorem 4, and q′0, q′1 and γ without size constraint. For λ

small enough, the corresponding (qλ, (qλ)′)(0) and (qλ, (qλ)′)(T ) satisfy the assumptions

of Theorem 4. Hence there exists a controlled solution (q, g) on [0, λT ], achieving

(q, q′)(0) = (q0, q′0) and (q, q′)(λT ) = (q1, q

′1), for λ small enough. Moreover we can

assume that it is the case without loss of generality that λ is small, and in particular

that λ ≤ 1. Thus the result is obtained but in a shorter time interval.

To get to the desired time interval, using that Equation (2.20) enjoys some invari-

ance properties by translation and time-reversal (up to the change of the sign of γ) it

is sufficient to glue together an odd number, say 2N + 1 with N in N∗, of appropriate

controlled solutions each defined on a time interval of length λT with λ = 12N+1 , going

back and forth between (q0, q′0) and (q1, q

′1) until time T = (2N + 1)λT . Moreover one

can see that the gluings are not only C2 but even C∞.

We have therefore already proven that Theorem 1 is true in the case where q1 is

close to q0, or more precisely for any q0 in Qδ and q1 in B(q0, rq0).

16

For the general case where q0 and q1 are in the same connected component of Qδ for

some δ > 0, without the closeness condition, we use again a gluing process. Consider

indeed a smooth curve from q0 to q1. For each point q on this curve, there is a rq > 0

such that for any q in B(q, rq), any q′0, q′1 and any γ, one can connect (q, q′0) to (q, q′1) by

a solution of the system, for any time T > 0. Extract a finite subcover of the curve by

the balls B(q, rq). Therefore we find N ≥ 2 and (q iN

)i=1,...,N−1 in the same connected

component of Qδ as q0 such that for any i = 1, . . . , N , q iN

is in B(q i−1N, rq i−1

N

) (note

that this includes q0 and q1). Therefore, using again the local result obtained above,

there exist some controlled solutions from (q i−1N, 0) to (q i

N, 0) (for i = 1 and i = N

we use (q0,q′0N ) and (q1,

q′1N ) rather than (q0, 0) and (q1, 0)), each on a time interval of

length T associated with circulation γN . One deduces by time-rescaling some controlled

solutions associated with circulation γ on a time interval of length TN . Gluing them

together leads to the desired controlled solution.

4 Reduction to an approximate controllability result

The goal of this section is to prove that Theorem 4 can be deduced from the following

approximate controllability result thanks to a topological argument already used in

[22], see Lemma 3 below. Let us mention that a similar argument has also been used

for control purposes but in other contexts, see e.g. [1, 5, 26, 27].

Theorem 5. Given δ > 0, S0 ⊂ Ω bounded, closed, simply connected with smooth

boundary, which is not a disk, q0 in Qδ and T > 0, there is r > 0 such that B(q0, r)

is included in the same connected component of Qδ as q0 and such that for any γ ∈ Rwith |γ| ≤ r and for any q′0 in B(0, r), for any η > 0, there is a mapping

T : B((q0, q

′0), r

)→ C∞([0, T ];Qδ)

which to (q1, q′1) associates q where (q, g) is a controlled solution associated with the

initial data (q0, q′0), such that the mapping

(q1, q′1) ∈ B

((q0, q

′0), r

)7→(T (q1, q

′1), T (q1, q

′1)′)(T ) ∈ Qδ × R3

is continuous and such that for any (q1, q′1) in B

((q0, q

′0), r

),

‖(T (q1, q

′1), T (q1, q

′1)′)(T )− (q1, q

′1)‖ 6 η.

The proof of Theorem 5 will be given in Section 5. Here we prove that Theorem 4

follows from Theorem 5.

Proof of Theorem 4 from Theorem 5. Let δ > 0, S0 ⊂ Ω bounded, closed, simply con-

nected with smooth boundary, which is not a disk, q0 in Qδ and T > 0. Let r > 0

17

as in Theorem 5. Let γ ∈ R with |γ| ≤ r and q′0 in B(0, r). According to Theorem 5

applied with η = r2 there is a mapping T : B

((q0, q

′0), r

)→ C∞([0, T ];Qδ) which maps

(q1, q′1) to q where (q, g) is a controlled solution associated with the initial data (q0, q

′0),

such that for any (q1, q′1) in B

((q0, q

′0), r

), ‖(T (q1, q

′1), T (q1, q

′1)′)(T ) − (q1, q

′1)‖ 6 r

2 .

We define a mapping f from B((q0, q

′0), r

)to R6 which maps (q1, q

′1) to f(q1, q

′1) :=(

T (q1, q′1), T (q1, q

′1)′)(T ). Then we apply the following lemma borrowed from [22, pages

32-33], to w0 = (q0, q′0) and κ = r.

Lemma 3. Let w0 ∈ Rn, κ > 0, f : B(w0, κ) → Rn a continuous map such that we

have |f(w)− w| ≤ κ2 for any x in ∂B(w0, κ). Then B(w0,

κ2 ) ⊂ f(B(w0, κ)).

This allows to conclude the proof of Theorem 4 setting r = κ2 = r

2 .

5 Proof of the approximate controllability result Theo-

rem 5

In this section we prove Theorem 5 by exploiting the geodesic feature of the uncontrolled

system with zero circulation, cf. the observation below Theorem 3. To do so, we will

use some well-chosen impulsive controls which allow to modify the velocity q′ in a short

time interval and put the state of the system on a prescribed geodesic (and use that

|γ| is small). We mention here [4] and the references therein for many more examples

on the impulsive control strategy.

5.1 First step

We consider S0 ⊂ Ω as before and consider δ > 0 so that q0 ∈ Qδ. We let r1 > 0 be

small enough so that B(q0, r1) ⊂ Qδ. We also let T > 0.

The first step consists in considering the geodesics associated to the uncontrolled,

potential case (γ = 0). The following classical result regarding the existence of geodesics

can be found for instance in [33, Section 7.5], see also [12] for the continuity feature.

Lemma 4. There exists r2 in (0, 12r1) such that for any q1 in B(q0, r2) there exists a

unique C∞ solution q(t) lying in B(q0,12r1) to(

Mg +Ma(q))q′′ + 〈Γ(q), q′, q′〉 = 0 on [0, T ], with q(0) = q0, q(T ) = q1. (5.1)

Furthermore the map q1 ∈ B(q0, r2) 7→ (c0, c1) ∈ R6 given by c0 = q′(0), c1 = q′(T ) is

continuous.

Let us fix r2 as in the lemma before. Let q′0 in B(0, r2) and (q1, q′1) in B

((q0, q

′0), r2

).

18

Our goal is to make the system follow approximately such a geodesic q which we

consider fixed during this Section. For the geodesic equation in (5.1), q0 and q1 deter-

mine the initial and final velocities (which of course differ in general from q′0 and q′1).

But we will see that is possible to use the penultimate term of (2.20) in order to modify

the initial and final velocities of the system. Precisely, the control will be used so that

the right hand side of (2.20) behaves like two Dirac masses at time close to 0 and T ,

driving the velocity q′ from the initial and final velocities to the ones of the geodesic in

two short time intervals close to 0 and T .

5.2 Illustration of the method on a toy model

Let us illustrate this strategy on a toy model. We will later on adapt the analysis to

the complete model, cf. Proposition 4.

Let β : R → R be a smooth, non-negative function supported in [−1, 1], such

that∫ 1−1 β(t)2 dt = 1 and, for ε in (0, 1), βε(t) := 1√

εβ(t−εε

), so that2 (β2

ε )ε is an

approximation of the unity when ε→ 0+.

For a function f defined on [0, T ], we will denote

‖f‖T,ε := ‖f‖C0([0,T ]) + ‖f‖C1([2ε,T−2ε]). (5.2)

Lemma 5. Let q0, r2, q1, q′0 and q′1 as above. Let

v0 :=(Mg +Ma(q0)

)(c0(q1)− q′0) and v1 := −

(Mg +Ma(q1)

)(c1(q1)− q′1). (5.3)

Let, for ε in (0, 1), qε the maximal solution to the following Cauchy problem:(Mg +Ma(qε)

)q′′ε + 〈Γ(qε), q

′ε, q′ε〉 = β2

ε (·) v0 + β2ε (T − ·)v1, (5.4)

with qε(0) = q0 and q′ε(0) = q′0. Then for ε small enough, qε(t) lies in B(q0, r1) for t in

[0, T ] and, as ε→ 0+, ‖qε − q‖T,ε → 0 and (qε, q′ε)(T )→ (q1, q

′1).

Proof. For ε in (0, 1), let us denote Tε = sup T > 0 such that qε(t) ∈ B(q0, r1) for t ∈(0, T ). Let us first prove that there exists T > 0 such that for any ε in (0, 1), Tε ≥ T .

Using the identity (2.15), we obtain indeed, for any ε in (0, 1), for any t ∈ (0, Tε),(Mg+Ma(qε(t))

)q′ε(t)·q′ε(t) =

(Mg+Ma(q0)

)q′0 ·q′0 +2

∫ t

0

(β2ε (·) v0 +β2

ε (T−·)v1

)·q′ε,

Moreover, relying on Remark 2, we see that there exists c > 0 (which depends on δ)

such that for any q in Qδ, for any p in R3,

c|p|2 ≤(Mg +Ma(q)

)p · p ≤ c−1|p|2. (5.5)

2In the next lemma we are going to make use only of the square function β2ε but we will also have

to deal with the function βε itself in the sequel, see below Proposition 2.

19

Therefore using Gronwall’s lemma we obtain that there exists C > 0 such that for any

ε in (0, 1), for any t ∈ (0, Tε), supt∈(0,Tε) ‖q′ε(t)‖ ≤ C. Therefore by the mean value

theorem for T := r1/2C, one has for any ε in (0, 1), Tε ≥ T .

We now prove in the same time that for ε > 0 small enough, Tε ≥ T , and the

convergence results stated in Lemma 5. In order to exploit the supports of the functions

βε(·) and βε(T−·) in the right hand side of the equation (5.4) we compare the dynamics

of qε and q during the three time intervals [0, 2ε], [2ε, T − 2ε] and [T − 2ε, T ].

For ε1 := T /2 and ε in (0, ε1), one already has that Tε ≥ 2ε and we can therefore

simply compare the dynamics of qε and q on the first interval [0, 2ε]. Indeed using again

the mean value theorem we obtain that supt∈[0,2ε] |qε − q0| converges to 0 as ε goes to

0. Moreover integrating the equation (5.4) on [0, 2ε] and taking into account the choice

of v0 in (5.3), we obtain(Mg +Ma(qε(2ε))

)q′ε(2ε) =

(Mg +Ma(q0)

)c0(q1)

−∫ 2ε

0

(DMa(qε) · q′ε

)· q′ε dt−

∫ 2ε

0〈Γ(qε), q

′ε, q′ε〉 dt, (5.6)

Now, there exists C > 0 such that for any q in Qδ, for any p in R3,

|(DMa(q) · p

)· p|+ |〈Γ(q), p, p〉| ≤ C|p|2. (5.7)

Combining this and the bound on q′ε we see that the two terms of the last line of (5.6)

above converge to 0 as ε goes to 0. Since q 7→ Ma(q) is continuous on Qδ and qε(2ε)

converges to q0 as ε→ 0, the matrixMa(qε) converges toMa(q0) as ε→ 0. Therefore,

using that the matrix Mg +Ma(q0) is invertible we deduce that q′ε(2ε) converges to

c0(q1) as ε goes to 0.

During the time interval [2ε, T − 2ε], the right hand side of the equation (5.4)

vanishes and the equation therefore reduces to the geodesic equation in (5.1). Since

this equation is invariant by translation in time, one may use the following elementary

result on the continuous dependence on the data, with a time shift of 2ε.

Lemma 6. There exists η > 0 such that for any (q0, q′0) in B((q0, c0(q1)), η) there exists

a unique C∞ solution q(t) lying in B(q0, r1) to(Mg +Ma(q)

)q′′ + 〈Γ(q), q′, q′〉 = 0

on [0, T ], with q(0) = q0, q′(0) = q′0. Furthermore ‖q − q‖C1([0,T ]) → 0 as (q0, q

′0) →

(q0, c0(q1)).

Since qε(2ε) and q′ε(2ε) respectively converge to q0 and c0(q1), according to Lemma 6

there exists ε2 in (0, ε1) such that for ε in (0, ε2), there exists a unique C∞ solution

qε(t) lying in B(q0, r1) to(Mg +Ma(qε)

)q′′ε + 〈Γ(qε), q

′ε, q′ε〉 = 0 on [0, T ], with qε(0) =

qε(2ε), q′ε(0) = q′ε(2ε) and ‖qε − q‖C1([0,T ]) → 0 as ε→ 0.

Since the function defined by qε(t) = qε(t + 2ε) also satisfies(Mg +Ma(qε)

)q′′ε +

〈Γ(qε), q′ε, q′ε〉 = 0 on [0, T − 4ε], with qε(0) = qε(2ε), q

′ε(0) = q′ε(2ε), by the uniqueness

20

part in the Cauchy-Lipschitz theorem one has that Tε ≥ T − 2ε and qε and qε coincide

on [0, T − 4ε], so that, shifting back in time, ‖qε − q(· − 2ε)‖C1([2ε,T−2ε]) → 0 as ε→ 0.

Since q is smooth, this entails that ‖qε − q‖C1([2ε,T−2ε]) → 0 as ε→ 0.

Finally one deals with the time interval [T −2ε, T ] in the same way as the first step.

In particular, reducing ε one more time if necessary one obtains, by an energy estimate,

a Gronwall estimate and the mean value theorem, that Tε ≥ T . Moreover the choice

of the vector v1 in (5.3) allows to reorientate the velocity q′ε from c1(q1) to q′1 whereas

the position is not much changed (due to the uniform bound of q′ε and the mean value

theorem) so that the value of qε at time T converges to q1 as ε goes to 0.

5.3 Back to the complete model

Now in order to mimic the right hand side of (5.4) we are going to use one part of the

force term F1 introduced in Definition 2. Let us therefore introduce some notations for

the different contributions of the force term F1. We define, for any q in Q, p in R3, α

in C∞(F(q);R),

F1,a(q)[α] := −1

2

∫∂S(q)

|∇α|2 ∂nΦ(q, ·) dσ, (5.8)

F1,b(q, p)[α] := −∫∂S(q)

∇α · ∇(p · Φ(q, ·)) ∂nΦ(q, ·) dσ, (5.9)

F1,c(q)[α] := −∫∂S(q)

∇α · ∇⊥ψ(q, ·) ∂nΦ(q, ·) dσ, (5.10)

so that for any γ in R,

F1(q, p, γ)[α] = F1,a(q)[α] + F1,b(q, p)[α] + γF1,c(q)[α].

The part which will allow us to approximate the right hand side of (5.4) is F1,a. More

precisely we are going to see (cf. Proposition 3) that there exists a control α (chosen

below as α = A[q, gε] with gε given by (5.14)) such that in the appropriate regime the

dynamics of (2.20) behaves like the equation with only F1,a on the right hand side.

Moreover the following lemma, where the time parameter does not appear, proves that

the operator F1,a(q)[·] can actually attain any value v in R3. Recall that δ > 0 has

been fixed at the beginning of Section 5.1.

Proposition 2. There exists a continuous mapping g : Qδ ×R3 → C such that for any

(q, v) in Qδ × R3 the function α := A[q, g(q, v)] in C∞(F(q);R) satisfies:

∆α = 0 in F(q), and ∂nα = 0 on ∂F(q) \ Σ, (5.11)∫∂S(q)

|∇α|2 ∂nΦ(q, ·) dσ = v, (5.12)∫∂S(q)

α∂nΦ(q, ·) dσ = 0. (5.13)

21

We recall that the operator A was introduced in Definition 1. The result above

will be proved in Section 7. Note that when S(q) is a homogeneous disk, an adapted

version of Proposition 2 still holds, see Proposition 7 in Section 7. The condition (5.13)

will be useful to cancel out the last term of (2.20).

We define

gε(t, x) := βε(t)g(q0,−2v0)(x) + βε(T − t)g(q1,−2v1)(x), (5.14)

where v0 and v1 defined in (5.3), for (q1, q′1) in B

((q0, q

′0), r2

), and g is given by Propo-

sition 2. The goal is to prove that for ε and |γ| small enough, this control drives the

system (2.20) with α = A[q, gε] from (q0, q′0) to (q1, q

′1), approximately.

1. We first observe that

F1,a(q)[A[q, gε]] = β2ε (t)F1,a(q)

[A[q, g(q0,−2v0)]

]+ β2

ε (T − t)F1,a(q)[A[q, g(q1,−2v1)]

], (5.15)

and is therefore a good candidate to approximate the right hand side of (5.4) if q is

near q0 for t near 0 and if q is near q1 for t near T . One then may indeed expect that

F1,a(q)[A[q, g(q0,−2v0)]

]and F1,a(q)

[A[q, g(q1,−2v1)]

]are close to

F1,a(q0)[A[q0, g(q0,−2v0)]

]and F1,a(q1)

[A[q1, g(q1,−2v1)]

], respectively,

on the respective supports of βε(·) and βε(T −·). Moreover, according to Proposition 2

these last two terms are equal to v0 and v1 (see (5.8) and (5.12)).

2. Next we will rigorously prove in Proposition 4 below that the conclusion of Lemma 5

for the toy system also holds when one substitutes the term F1,a(q)[A[q, gε]] in (5.15).

This corresponds also to (2.20) with γ = 0 and the term F1,b and F2 put to zero.

3. Finally it will appear that in an appropriate regime, in particular for small ε and

|γ|, the second last term of (2.20) is dominant with respect to the other terms of the

right hand side (here the condition (5.13) above will be essential in order to deal with

the last term of (2.20)).

Let us state a proposition summarizing the claims above. According to the Cauchy-

Lipschitz theorem there exists a controlled solution qε,γ associated with the control gε

introduced in (5.14), starting with the initial condition qε,γ(0) = q0 and q′ε,γ(0) = q′0,

22

with circulation γ, and lying in B(q0, r1) up to some positive time Tε,γ . More explicitly

qε,γ satisfies on [0, Tε,γ ],(Mg +Ma(qε,γ)

)q′′ε,γ + 〈Γ(qε,γ), q′ε,γ , q

′ε,γ〉 = γ2E(qε,γ) + γq′ε,γ ×B(qε,γ)

+ F1(qε,γ , q′ε,γ , γ)

[A[qε,γ , gε]

]+ F2(qε,γ)

[∂tA[qε,γ , gε]

]. (5.16)

Observe that due to the choice of the control gε in (5.14) the function qε,γ also depends

on (q1, q′1) through v0 and v1, see their definition in (5.3).

We have the following approximation result.

Proposition 3. For ε and |γ| small enough, Tε,γ > T and, as ε and |γ| converge to 0+,

‖qε,γ − q‖T,ε → 0 and (qε,γ , q′ε,γ)(T )→ (q1, q

′1), uniformly for (q1, q

′1) in B

((q0, q

′0), r2

).

This result will be proved in Section 6. Once Proposition 3 is proved, Theorem 5

follows rapidly. Indeed, according to this proposition, for η > 0, there exists ε small

enough and r in (0, r2) such that for any γ ∈ R with |γ| ≤ r and for any q′0 in

B(0, r), the mapping T defined on B((q0, q

′0), r

)by setting T (q1, q

′1) = qε,γ , has the

desired properties. In particular the continuity of T follows from the regularity of c0

in Lemma 4 and of the solution of ODEs on their initial data. This ends the proof of

Theorem 5.

5.4 About Remark 1

Now that we presented the scheme of proof of Theorem 1 let us explain how to obtain

the improvement mentioned in Remark 1. It is actually a direct consequence of the

explicit formula for gε(t, x) given in (5.14) and of a change of variable in time. Due to

the expression of βε given at the beginning of Section 5.2 one obtains that the total

flux through Σ−, that is∫ T

0

∫Σ− gε dσdt, is of order

√ε. Hence one can reduce ε again

in order to satisfy the requirement of Remark 1.

On the other hand observe that the time-rescaling argument used in the proof of

Theorem 1 from Theorem 4, cf. (3.2), leaves the total flux through Σ− invariant, while

the number N of steps involved in the end of the same proof does not depend on ε.

6 Closeness of the controlled system to the geodesic. Proof

of Proposition 3

In this section, we prove Proposition 3.

6.1 Proof of Proposition 3

The proof of Proposition 3 is split in several parts. To compare qε,γ and q, we are going

to consider an “intermediate trajectory” qε which imitates the trajectory qε of the toy

23

model of Lemma 5, by using the part F1,a of the force term. More precisely we define

qε by(Mg +Ma(qε)

)q′′ε + 〈Γ(qε), q

′ε, q′ε〉 = F1,a(qε)

[A[qε, gε]

],

with qε(0) = q0, q′ε(0) = q′0, (6.1)

where gε was defined in (5.14) and where the operator A was introduced in Defini-

tion 1. Note that due to the definition of gε, the function qε also depends on q1, q′1.

The statement below is an equivalent of Lemma 5 for qε, comparing qε to the “target

geodesic” q.

Proposition 4. There exists ε1 > 0 such that, for any ε ∈ (0, ε1], for any (q1, q′1)

in B((q0, q

′0), r2

), the solution qε given by (6.1) lies in the ball B(q0, r1) at least up to

T . Moreover ‖qε − q‖T,ε converges to 0 and (qε, q′ε)(T ) converges to (q1, q

′1) when ε

converges to 0+, uniformly for (q1, q′1) in B

((q0, q

′0), r2

)for both convergences.

We recall that the norm ‖ · ‖T,ε was defined in (5.2). The proof of Proposition 4

can be found in Subsection 6.2.

The following result allows us to deduce the closeness of the trajectories qε,0, given

by (5.16) with γ = 0, and qε given by (6.1). Let us recall that by the definition of Tε,γ

that comes along (5.16), qε,0 lies in B(q0, r1) up to the time Tε,0, which depends on

q1, q′1.

Proposition 5. There exists ε2 in (0, ε1] such that for any ε ∈ (0, ε2], one has Tε,0 ≥ T .

Moreover ‖qε− qε,0‖C1([0,T ]) → 0 when ε→ 0+, uniformly for (q1, q′1) in B

((q0, q

′0), r2

).

The proof of Proposition 5 can be found in Subsection 6.3.

Finally, we have the following estimation of the deviation due to the circulation γ,

which will be proved in Subsection 6.4.

Proposition 6. There exists ε3 in (0, ε2] such that for all ε ∈ (0, ε3], there exists γ0 > 0

such that for any γ ∈ [−γ0, γ0], we have Tε,γ ≥ T and ‖qε,γ − qε,0‖C1[0,T ] converges to

0 when γ → 0, uniformly for (q1, q′1) in B

((q0, q

′0), r2

).

Propositions 4, 5 and 6 give us directly the result of Proposition 3.

6.2 Proof of Proposition 4

We proceed as in the proof of Lemma 5 with a few extra complications related to the

fact that the right hand side of the equation (6.1) is more involved than the one of

the equation (5.4) and to the fact that we need to obtain uniform convergences with

respect to (q1, q′1) in B

((q0, q

′0), r2

).

24

As in the proof of Lemma 5 we introduce, for ε in (0, 1), the time Tε = sup T >

0 such that qε(t) ∈ B(q0, r1) for t ∈ (0, T ) and we first prove that there exists T > 0

such that for any ε in (0, 1), Tε ≥ T thanks to an energy estimate. In order to deal with

the term coming from (5.15) in the right hand side of the energy estimate, recalling

Remark 2 and the definition of F1,a in (5.8), we observe that for any R > 0, there exists

C > 0 such that for any q, q in Qδ, for any v in B(0, R),

|F1,a(q)[A[q, g(q, v)]

]| ≤ C. (6.2)

This allows to deduce from the expressions of v0 and v1 in (5.3) that there exists T > 0

and C > 0 such that for any (q1, q′1) in B

((q0, q

′0), r2

), for any ε in (0, 1), Tε ≥ T and

‖q′ε‖C([0,Tε]) ≤ C. We deduce that for ε1 := T /2 and ε in (0, ε1), Tε ≥ 2ε and that

supt∈[0,2ε] |qε − q0| converges to 0 as ε goes to 0 uniformly in (q1, q′1) in B

((q0, q

′0), r2

).

Now let us prove that q′ε(2ε) converges to c0(q1) as ε goes to 0 uniformly in (q1, q′1)

in B((q0, q

′0), r2

). We integrate the equation (6.1) on [0, 2ε]. Thus(

Mg +Ma(qε(2ε)))q′ε(2ε) =

(Mg +Ma(q0)

)q′0

−∫ 2ε

0

(DMa(qε) · q′ε

)· q′ε dt−

∫ 2ε

0〈Γ(qε), q

′ε, q′ε〉 dt+

∫ 2ε

0F1,a(qε)

[A[qε, gε]

]dt. (6.3)

Then we pass to the limit as ε goes to 0+ in the last equality. Here we use two extra

arguments with respect to the corresponding argument in the proof of Lemma 5. On

the one hand we see that the convergences ofMa(qε(2ε)) toMa(q0) and of the two first

terms of the last line to 0, already obtained in the proof of Lemma 5, hold uniformly

with respect to (q1, q′1) in B

((q0, q

′0), r2

), as a consequence of the uniform estimates of

qε−q0 and q′ε obtained above. On the other hand the term F1,a enjoys the following reg-

ularity property with respect to q: we have that q 7→ F1,a(q)[A[q, g(q0, v)]

]is Lipschitz

with respect to q in Qδ uniformly for v in bounded sets of R3. Therefore using that

supt∈[0,2ε] |qε−q0| converges to 0 as ε goes to 0 uniformly in (q1, q′1) in B

((q0, q

′0), r2

), the

expressions of v0 and v1 in (5.3) and that F1,a(q0)[A[q0, g(q0,−2v0)

]= v0, according

to Proposition 2 we deduce that

supt∈[0,2ε]

∣∣∣F1,a(qε)[A[qε, g(q0,−2v0)]

]− v0

∣∣∣converges to 0 as ε goes to 0 uniformly in (q1, q

′1) in B

((q0, q

′0), r2

). Since for t in [0, 2ε],

the equation (5.15) applied to q = qε is simplified into

F1,a(qε)[A[qε, gε]] = β2ε (t)F1,a(qε)

[A[qε, g(q0,−2v0)]

],

and that∫ 2ε

0 β2ε (t) dt = 1, we get that the last term in (6.3) converges to v0 when ε

goes to 0. Moreover, due to the choice of v0 the first and last term of the right hand

side of (6.3) can be combined at the limit to get(Mg +Ma(q0)

)c0(q1).

25

Therefore, inverting the matrix in the right hand side of (6.3) and passing to the

limit, we see that q′ε(2ε) converges to c0(q1) as ε goes to 0 uniformly in (q1, q′1) in

B((q0, q

′0), r2

).

When t is in [2ε, T − 2ε], the equation (6.1) reduces to a geodesic equation so that

the same arguments as in the proof of Lemma 5 apply.

Finally for the last step, for t in [T − 2ε, T ], we proceed in the same way as in the

first step. This ends the proof of Proposition 4.

6.3 Proof of Proposition 5

We begin with the following lemma, which provides a uniform boundedness for the

trajectories qε,0 satisfying (5.16) with γ = 0, that is(Mg +Ma(qε,0)

)q′′ε,0 + 〈Γ(qε,0), q′ε,0, q

′ε,0〉 = F1,a(qε,0)

[A[qε,0, gε]

]+ F1,b(qε,0, q

′ε,0)[A[qε,0, gε]

]+ F2(qε,0)

[∂tA[qε,0, gε]

]. (6.4)

We recall that gε is given by (5.14) with v0 and v1 given by (5.3). The terms F1,a and

F1,b were defined in (5.8)-(5.9), F2 in (2.18). Also we recall that by definition of Tε,0

(see the definition of Tε,γ in the end of Subsection 5.3), during the time interval [0, Tε,0],

qε,0 remains in B(q0, r1).

Lemma 7. There exists εa > 0 such that

sup(q1,q′1)∈B((q0,q′0),r2),

ε∈(0,εa]

‖q′ε,0‖C([0,Tε,0]) < +∞.

Proof. First we see that the mappings

q 7→ F1,a(q)[A[q, g(q0, v)]] and q 7→ F1,b(q, ·)[A[q, g(q0, v)]]

are bounded for q in Qδ, uniformly for v in bounded sets of R3. Let us now focus on

the F2 term. For t in [0, 2ε], gε(t) = βε(t)g(q0,−2v0) so that, by the chain rule, for t in

[0,min(2ε, Tε,0)],

∂tA[qε,0, gε] = βεDqA[qε,0, g(q0,−2v0)] · q′ε,0 + β′εA[qε,0, g(q0,−2v0)].

For what concerns F2 we have, using the property (5.13),

F2(qε,0)[∂tA[qε,0, gε]

]= βε

∫∂S(qε,0)

(DqA[qε,0, g(q0,−2v0)] · q′ε,0

)∂nΦ(qε,0, ·) dσ

+β′ε

(∫∂S(qε,0)

A[qε,0, g(q0,−2v0)] ∂nΦ(qε,0, ·) dσ−∫∂S(q0)

A[q0, g(q0,−2v0)] ∂nΦ(q0, ·) dσ).

Using that the mapping q 7→∫∂S(q)∇qA[q, g(q0, v)] ⊗ ∂nΦ(q, ·) dσ is bounded for

q over Qδ and that the mapping q 7→∫∂S(q)A[q, g(q0, v)] ∂nΦ(q, ·) dσ is Lipschitz with

26

respect to q in Qδ, both uniformly for v in bounded sets of R3, we see that this involves

(recalling the expression of βε given at the beginning of Section 5.2)

∣∣F2(qε,0)[∂tA[qε,0, gε]

]∣∣ . C

(1

ε1/2|q′ε,0|+

1

ε3/2|qε,0 − q0|

), (6.5)

uniformly for (q1, q′1) in B

((q0, q

′0), r2

). Then, multiplying (6.4) by q′ε,0 and using once

more the identity (2.15), we obtain, for any ε in (0, 1), for t in [0,min(2ε, Tε,0)],(Mg +Ma(qε,0(t))

)q′ε,0(t) · q′ε,0(t) =

(Mg +Ma(q0)

)q′0 · q′0

+2

∫ t

0

(F1,a(qε,0)

[A[qε,0, gε]

]+F1,b(qε,0, q

′ε,0)[A[qε,0, gε]

]+F2(qε,0)

[∂tA[qε,0, gε]

])·q′ε,0,

(6.6)

Then, using (5.5), the boundedness of the mappings q 7→ F1,a(q)[A[q, g(q0, v)]] and

q 7→ F1,b(q, ·)[A[q, g(q0, v)]] already mentioned above, the definition of βε and the bound

(6.5), we get

|q′ε,0(t)|2 ≤ C(

1 +1

ε1/2

∫ t

0|q′ε,0(s)|2 ds+

1

ε3/2

∫ t

0|q′ε,0(s)||qε,0(s)− q0| ds

).

Then using the mean value theorem and that t ≤ 2ε, we have that

|q′ε,0(t)|2 ≤ C

(1 + ε1/2 sup

[0,min(2ε,Tε,0)]|q′ε,0|2

),

so that for ε small enough, and for t in [0,min(2ε, Tε,0)], |q′ε,0(t)| ≤ C, uniformly for

(q1, q′1) in B

((q0, q

′0), r2

). As a consequence of the usual blow-up criterion for ODEs,

we have that Tε,0 ≥ 2ε.

During the next phase, i.e. for t in [2ε, T − 2ε], the control is inactive so that the

equation (6.4) is a geodesic equation. Then by a simple energy estimate we get again

that |q′ε,0(t)| ≤ C on [0,min(T − 2ε, Tε,0)].

Finally if Tε,0 ≥ T − 2ε, then we deal with the last phase as in the first phase. This

concludes the proof of Lemma 7.

We then conclude the proof of Proposition 5 by a classical comparison argument

using Gronwall’s lemma and the Lipschitz regularity with respect to q of the various

mappings involved (Ma, Γ, F1,a, F1,b and F2). This allows to prove that there exists

ε2 in (0, ε1] such that for any ε ∈ (0, ε2], Tε,0 ≥ T and ‖qε − qε,0‖C1([0,T ]) → 0 when

ε→ 0+, uniformly for (q1, q′1) in B

((q0, q

′0), r2

). This ends the proof of Proposition 5.

6.4 Proof of Proposition 6

First we may extend Lemma 7 to the solutions qε,γ to (5.16) in the following manner.

27

Lemma 8. There exists εb in (0, ε2) such that ‖q′ε,γ‖C([0,Tε,γ ]) is bounded uniformly in

ε ∈ (0, εb], for any γ ∈ [−1, 1], and for (q1, q′1) ∈ B((q0, q

′0), r2).

It is indeed a matter of adding the “electric field” E in (6.6), and noting that E is

bounded on Qδ; the “magnetic field” B gives no contribution to the energy.

We now finish the proof of Proposition 6. Using a comparison argument we obtain

that there exists ε3 in (0, εb] such that for all ε ∈ (0, ε3], there exists γ0 > 0 such that for

any γ ∈ [−γ0, γ0], we have Tε,γ ≥ T and ‖qε,γ − qε,0‖C1[0,T ] converges to 0 when γ → 0,

uniformly for (q1, q′1) in B

((q0, q

′0), r2

). This concludes the proof of Proposition 6.

7 Design of the control according to the solid position.

Proof of Proposition 2

This section is devoted to the proof of Proposition 2.

7.1 The case of a homogeneous disk

Before proving Proposition 2 we establish the following similar result concerning the

simpler case where the solid is a homogeneous disk. In that case, the statement merely

considers q of the form q = (h, 0). Thus in order to simplify the writing, we introduce

Qhδ := h ∈ R2 such that (h, 0) ∈ Qδ.

Also in all this section when we will write q, it will be understood that q is associated

with h by q = (h, 0).

Proposition 7. Let δ > 0. Then there exists a continuous mapping g : Qhδ × R2 → Csuch that the function α := A[q, g(q, v)] in C∞(F(q);R) satisfies:

∆α(q, x) = 0 in F(q), and ∂nα(q, x) = 0 on ∂F(q) \ Σ, (7.1)∫∂S(q)

|∇α(q, x)|2 ndσ = v, (7.2)∫∂S(q)

α(q, x)ndσ = 0. (7.3)

In order to prove Proposition 7, the mapping g will be constructed using a combi-

nation of some elementary functions which we introduce in several lemmas.

To begin with, we will make use of the elementary geometrical property that

n(q0, x) : x ∈ ∂S(q0) is the unit circle S1 and of the following lemma.

28

Lemma 9. There exist three vectors e1, e2, e3 ∈ n(q0, x) : x ∈ ∂S(q0) and positive

C∞ maps (µi)16i63 : R2 → R+ such that for any v ∈ R2,

3∑i=1

µi(v)ei = v. (7.4)

Proof. One may consider for instance e1 := (1, 0), e2 := (0, 1), e3 := (−1,−1), and

µ1(v) = v1 +√

1 + |v1|2 + |v2|2, µ2(v) = v2 +√

1 + |v1|2 + |v2|2

and µ3(v) =√

1 + |v1|2 + |v2|2.

In the next lemma, we introduce some functions that are defined in a neighbourhood

of ∂S(q0) (for some q0 = (h0, 0) fixed), satisfying some counterparts of the properties

(7.1) and (7.2).

Lemma 10. There exist families of functions (αi,jε )ε∈(0,1), i, j ∈ 1, 2, 3, such that

for any i, j ∈ 1, 2, 3, for any ε ∈ (0, 1), αi,jε is defined and harmonic in a closed

neighbourhood V i,jε of ∂S(q0), satisfies ∂nαi,jε = 0 on ∂S(q0), and moreover one has for

any i, j, k, l in 1, 2, 3,∫∂S(q0)

∇αi,jε · ∇αk,lε ndσ → δ(i,j),(k,l) ei as ε→ 0+.

Proof. Without loss of generality, we may suppose that S(q0) is the unit disk. Consider

the parameterisation c(s) = (cos(s), sin(s)), s ∈ [0, 2π] of ∂S(q0) and the correspond-

ing si such that n(q0, c(si)) = ei, i ∈ 1, 2, 3.We consider families of smooth functions βi,jε : [0, 2π]→ R, i, j ∈ 1, 2, 3, ε ∈ (0, 1),

such that supp βi,jε ∩ supp βk,lε = ∅ whenever (i, j) 6= (k, l), diam(

supp βi,jε)→ 0 as

ε→ 0+,∫ 2π

0βi,jε (s) dσ = 0 and

∣∣∣∣∫ 2π

0|βi,jε (s)|2 n(q0, c(s))ds− ei

∣∣∣∣→ 0 as ε→ 0+.

Then we define αi,jε in polar coordinates as the truncated Laurent series:

αi,jε (r, θ) :=1

2

∑0<k≤K

1

k

(rk +

1

rk

)(−bi,jk,ε cos(kθ) + ai,jk,ε sin(kθ)),

where ai,jk,ε and bi,jk,ε denote the k-th Fourier coefficients of the function βi,jε . It is elemen-

tary to check that the function αi,jε satisfies the required properties for an appropriate

choice of K.

29

Now, for any h ∈ Qhδ , we may define V i,jε (q) := V i,jε −h0+h, which is a neighborhood

of ∂S(q), and αi,jε (q, x) := αi,jε (x + h0 − h), for each x ∈ V i,jε (q). We have for i, j, k, l

in 1, 2, 3,∫∂S(q)

∇αi,jε (q, x) · ∇αk,lε (q, x)n(q, x) dσ =

∫∂S(q0)

∇αi,jε (x) · ∇αk,lε (x)n(q0, x) dσ.

Proceeding as in [14] (see also [13, p. 147-149]) and relying in particular Runge’s theo-

rem, we have the following result which asserts the existence of harmonic approximate

extensions on the whole fluid domain.

Lemma 11. There exists a family of functions (αi,jη )η∈(0,1), i, j ∈ 1, 2, 3, harmonic

in F(q), satisfying ∂nαi,jη (q, x) = 0 on ∂F(q) \ Σ, with for any k in N,

‖αi,jη (q, ·)− αi,jε (q, ·)‖Ck(Vi,jε (q)∩F(q))

→ 0 when η → 0+. (7.5)

We now check that the above construction can be made continuous in q.

Lemma 12. For any ν > 0, there exist continuous mappings h ∈ Qhδ 7→ αi,j(q, ·) ∈C∞(F(q)) where q = (h, 0), i, j ∈ 1, 2, 3, such that for any h ∈ Qhδ , ∆xα

i,j(q, x) = 0

in F(q), ∂nαi,j(q, x) = 0 on ∂F(q) \ Σ and∣∣∣∣∣

∫∂S(q)

∇αi,j(q, ·) · ∇αk,l(q, ·)ndσ − δ(i,j),(k,l) ei

∣∣∣∣∣ ≤ ν. (7.6)

Proof. Let us assume that the functions αi,jη were previously defined not only for h ∈ Qhδbut for h ∈ Qhδ ; this is possible by using a smaller δ. Hence we may for each h ∈ Qhδfind functions αi,jη (for some η > 0) satisfying the properties above, and in particular

such that (7.6) is valid.

Next we observe that for any h ∈ Qhδ , setting q = (h, 0), the unique solution

αi,jη (q, q, ·) (up to an additive constant) to the Neumann problem ∆xαi,jη (q, q, x) = 0 in

F(q), ∂nαi,jη (q, q, x) = 0 on ∂F(q) \ Σ, ∂nα

i,jη (q, q, x) = ∂nα

i,jη (q, x) on Σ, is continuous

with respect to q ∈ Qδ. It follows that when a family of functions αi,jη satisfies (7.6) at

some point h ∈ Qhδ , it satisfies (7.6) (with perhaps 2ν in the right hand side) in some

neighborhood of h. Since Qhδ is compact and can be covered with such neighborhoods,

one can extract a finite subcover and use a partition of unity (according to the variable

q) adapted to this subcover to conclude: one gets an estimate like (7.6) with Cν on

the right hand side (for some constant C). It is then just a matter of considering ν/C

rather than ν at the beginning.

Finally our basic bricks to prove Proposition 7 are given in the following lemma,

where we can add the constraint (7.3).

30

Lemma 13. For any ν > 0, there exist continuous mappings q = (h, 0) ∈ Qδ 7→αi(q, ·) ∈ C∞(F(q)), i ∈ 1, 2, 3, such that for any q = (h, 0) ∈ Qδ, ∆xα

i(q, x) = 0 in

F(q), ∂nαi(q, x) = 0 on ∂F(q) \ Σ and∣∣∣∣∣

∫∂S(q)

∇αi(q, ·) · ∇αj(q, ·)ndσ − δi,j ei

∣∣∣∣∣ ≤ ν, (7.7)∫∂S(q)

αi(q, ·)ndσ = 0. (7.8)

Proof. Consider the functions αi,j given by Lemma 12. For any q = (h, 0) ∈ Qδ, for

any i ∈ 1, 2, 3, the three vectors∫∂S(q) α

i,j(q, ·)ndσ, where j ∈ 1, 2, 3, are linearly

dependent in R2; therefore there exists λi,j(q) ∈ R such that

3∑j=1

λi,j(q)

∫∂S(q)

αi,j(q, ·)ndσ = 0 and

3∑j=1

|λi,j(q)|2 = 1, (7.9)

Then one defines αi(q, ·) :=∑3

j=1 λi,j(q)αi,j(q, ·), and one checks that it satisfies (7.7)

with some Cν in the right hand side. Again changing ν in ν/C allows to conclude.

We are now in position to prove Proposition 7.

Proof of Proposition 7. Let δ > 0. Let ν > 0. We define the mapping S which to

(h, v) ∈ Qhδ × R2 associates the function

α(q, ·) :=3∑i=1

√µi(v)αi(q, ·),

in C∞(F(q)), where the functions µi were introduced in Lemma 9 and the functions αi

were introduced in Lemma 13. Next we define T : Qhδ × R2 → Qhδ × R2 by

(h, v) 7→ (T1, T2)(h, v) :=

(h,

∫∂S(q)

|∇α(q, ·)|2 ndσ

), where α = S(h, v).

Using (7.4) and (7.7), one checks that T is smooth and that

∂T2

∂v= Id +O(ν).

Hence taking ν sufficiently small, we see that ∂T2∂v is invertible, hence ∂T

∂(h,v) is invertible.

Consequently one can use the inverse function theorem on T : for each h0 ∈ Qhδ it realizes

a local diffeomorphism at(h0, 0), and hence on Qhδ × B(0, r) for r > 0 small enough.

This gives the result of Proposition 7 for v small: given (h, v) ∈ Qhδ × B(0, r), we let

(h, v) := T −1(h, v). Then the functions α :=∑3

i=1

√µi(v)αi(q, ·) and g := 1Σ ∂nα

satisfy the requirements. The general case follows by linearity of (7.1) and (7.3) and

by homogeneity of (7.2). This ends the proof of Proposition 7.

31

7.2 The case when S0 is not a disk

We now get back to the proof of Proposition 2. We will denote by coni(A) the conical

hull of A, namely

coni(A) :=

k∑i=1

λiai, k ∈ N∗, λi ≥ 0, ai ∈ A

,

The first step is the following elementary geometric lemma.

Lemma 14. Let S0 ⊂ Ω bounded, closed, simply connected with smooth boundary,

which is not a disk. Then coni(n(x), (x− h0)⊥ · n(x)), x ∈ ∂S0 = R3.

Proof. Suppose the contrary. Then there exists a plane separating (in the large sense)

the origin in R3 from the set coni((n(x), (x− h0)⊥ · n(x)), x ∈ ∂S0). We claim that

a normal vector to this plane can be put in the form (a, b, 1), with a, b ∈ R. Indeed,

otherwise it would need to be of the form (a, b, 0), and the separation inequality would

give (a, b) · n(x) ≥ 0, ∀x ∈ ∂S0. However, since ∂S0 is a smooth, closed curve, the set

n(x) : x ∈ ∂S0 is the unit circle of R2, therefore we have a contradiction.

Now we deduce that we have the following separation property:

(a, b) · n(x) + (x− h0)⊥ · n(x) ≥ 0, ∀x ∈ ∂S0.

Denoting w = (a, b) − h⊥0 , this translates into (w + x⊥) · n(x) ≥ 0. But using Green’s

formula, we get

0 ≤∫∂S0

(w + x⊥) · n(x) dσ =

∫S0

div(w + x⊥) dx = 0,

and consequently, we deduce that (w+x⊥) ·n(x) = 0 for all x in ∂S0. This is equivalent

to (x− w⊥) · τ(x) = 0 for all x in ∂S0. Parameterizing the translated curve ∂S0 − w⊥

by c(s), s ∈ [0, 1], it follows that c(s) · c(s) = 0, for all s in [0, 1], and therefore

|c(s)|2 is constant. This means that ∂S0 − w⊥ is a circle, so S0 is a disk, which is a

contradiction.

Fix q0 ∈ Qδ. Recalling the definitions of the Kirchhoff potentials in (2.2) and (2.3),

we infer from the previous lemma that

coni∂nΦ(q0, x), x ∈ ∂S0 = R3.

In place of Lemma 9, we have the following lemma which is a straightforward conse-

quence of Lemma 14 and of a repeated application of Caratheodory’s theorem on the

convex hull.

Lemma 15. There are some (xi)i∈1,...,16 in ∂S0 and positive continuous mappings

µi : R3 → R, 1 6 i 6 16, v 7→ µi(v) such that∑16

i=1 µi(v)∂nΦ(q0, xi) = v.

32

We are now in position to establish Proposition 2. We deduce from Lemma 15 that

for any q := (h, ϑ) ∈ Qδ, for any v in R3,

16∑i=1

µi(R(ϑ)v) ∂nΦ(q, xi(q)) = R(ϑ)v,

where xi(q) := R(ϑ)(xi − h0) + h and R(ϑ) denotes the 3 × 3 rotation matrix defined

by

R(ϑ) :=

(R(ϑ) 0

0 1

).

Due to the Riemann mapping theorem, there exists a biholomorphic mapping Ψ :

C\B(0, 1)→ C\S(q) with ∂S(q) = Ψ(∂B(0, 1)), where C denotes the Riemann sphere.

We consider the parametrisations c(s) = (cos(s), sin(s)), s ∈ [0, 2π] of ∂B(0, 1),

respectively Ψ(c(s)), s ∈ [0, 2π] of ∂S(q), and the corresponding si such that xi(q) =

Ψ(c(si)), for i ∈ 1, . . . , 16.Then, for any smooth function α : ∂S(q) → R, due to the Cauchy-Riemann rela-

tions, we have the following:

∂nα(Ψ(x)) =1√

|det(DΨ(x))|∂nB (α Ψ)(x),∫

∂S(q)|∇α(x)|2 ∂nΦ(q, x) dσ =

∫∂B(0,1)

|∇α(Ψ(x))|2 ∂nBΦ(q,Ψ(x))1√

|det(DΨ(x))|dσ,

for any x ∈ ∂B(0, 1), where n and nB respectively denote the normal vectors on ∂S(q)

and ∂B(0, 1). Note that, since Ψ is invertible, we have |det(DΨ(x))| > 0, for any

x ∈ ∂B(0, 1).

For each ε > 0, i ∈ 1, . . . , 16, j ∈ 1, 2, 3, 4 (here the index j belongs to 1, 2, 3, 4rather than 1, 2, 3 in order to adapt the linear dependence argument of Lemma 13

to the case of the three linear constraints (5.13)), we consider families of smooth

functions βi,jε : [0, 2π] → R satisfying supp βi,jε ∩ supp βk,lε = ∅ for (i, j) 6= (k, l),

diam(

supp βi,jε)→ 0 as ε→ 0+,

∫ 2π

0βi,jε (s) ds = 0,

and ∣∣∣∣∣∫ 2π

0|βi,jε (s)|2 ∂nΦ(q, c(s))

1√|det(DΨ(c(s)))|

ds− ei

∣∣∣∣∣→ 0 as ε→ 0+,

where

ei :=1√

|det(DΨ(c(si)))|∂nΦ(q, xi(q)).

Then one may proceed essentially as in the proof of Proposition 7. The details are

therefore left to the reader.

33

Acknowledgements

We would like to thank Jimmy Lamboley and Alexandre Munnier for helpful conver-

sations on shape differentiation. The authors also thank the Agence Nationale de la

Recherche, Project DYFICOLTI, grant ANR-13-BS01-0003-01 and Project IFSMACS,

grant ANR-15-CE40-0010 for their financial support. F. Sueur was also supported by

the Project BORDS, grant ANR-16-CE40-0027-01. Furthermore, J. J. Kolumban would

also like to thank the Fondation Sciences Mathematiques de Paris for their support in

the form of the PGSM Phd Fellowship.

References

[1] G. Aronsson, Global Controllability and Bang-Bang Steering of Certain Nonlinear

Systems, SIAM Journal on Control, 11 (1973), no. 4, 607–619.

[2] M. Boulakia, S. Guerrero, Local null controllability of a fluid-solid interaction

problem in dimension 3, J. European Math Society, 15 (2013), no. 3, 825–856.

[3] M. Boulakia, A. Osses, Local null controllability of a two-dimensional fluid-

structure interaction problem, ESAIM Control Optim. Calc. Var., 14 (2008), no.

1, 1–42.

[4] A. Bressan, Impulsive Control Systems, Nonsmooth Analysis and Geometric Meth-

ods in Deterministic Optimal Control, Volume 78 of the series The IMA Volumes

in Mathematics and its Applications, 1–22.

[5] P. Brunovsky, C. Lobry, Controlabilite Bang Bang, controlabilite differentiable,

et perturbation des systemes non lineaires. Ann. Mat. Pura Appl. (4) 105 (1975),

93–119.

[6] T. Chambrion, A. Munnier, Generic controllability of 3d swimmers in a perfect

fluid. SIAM Journal on Control and Optimization, 50(5) (2012), 2814–2835.

[7] J.-M. Coron, Exact boundary controllability of the Euler equations of incompress-

ible perfect fluids in dimension two, C. R. Acad. Sci. Paris, Serie I, 317 (1993),

271–276.

[8] J.-M. Coron, On the controllability of the 2-D incompressible Navier-Stokes equa-

tions with the Navier slip boundary conditions, ESAIM Controle Optim. Calc.

Var. 1 (1995/96), 35–75.

[9] J.-M. Coron, On the null asymptotic stabilization of two-dimensional incompress-

ible Euler equations in a simply connected domain, SIAM J. Control Optim., 37

(1999), 1874–1896.

34

[10] J.-M. Coron, F. Marbach, F. Sueur, Small time global exact null controllability

of the Navier-Stokes equation with Navier slip-with-friction boundary conditions.

Preprint 2016. http://arxiv.org/abs/1612.08087.

[11] J.-M. Coron, F. Marbach, F. Sueur, On the controllability of the Navier-

Stokes equation in spite of boundary layers. Proceeding of the RIMS confer-

ence “Mathematical Analysis of Viscous Incompressible Fluid”. Preprint 2017.

http://arxiv.org/abs/1703.07265.

[12] R. Gaines, Continuous dependence for two-point boundary value problems, Pacific

J. Math. 28 (1969), no. 2, 327–336.

[13] O. Glass, Some questions of control in fluid mechanics. Control of partial differen-

tial equations, 131–206, Lecture Notes in Math. 2048, Fond. CIME/CIME Found.

Subser., Springer, Heidelberg, 2012.

[14] O. Glass, An addendum to a J. M. Coron theorem concerning the controllability

of the Euler system for 2D incompressible inviscid fluids. J. Math. Pures Appl. (9)

80 (2001), no. 8, 845–877.

[15] O. Glass, Asymptotic stabilizability by stationary feedback of the 2-D Euler equa-

tion: the multiconnected case, SIAM J. Control Optim. 44 (2005), no. 3, 1105–

1147.

[16] O. Glass, Exact boundary controllability of the 3D Euler equation, ESAIM: Con-

trol, Optimization and Calculus of variations, 2000.

[17] O. Glass, T. Horsin, Prescribing the motion of a set of particles in a 3D perfect

fluid, SIAM J. Control Optim. 50 (2012), No. 5, 2726–2742.

[18] O. Glass, T. Horsin, Approximate Lagrangian controllability for the 2-D Euler

equation. Application to the control of the shape of vortex patches. J. Math. Pures

Appl. (9) 93 (2010), no. 1, 61–90.

[19] O. Glass, T. Horsin, Lagrangian controllability at low Reynolds number. ESAIM

Control Optim. Calc. Var. 22 (2016), no. 4, 1040–1053.

[20] O. Glass, C. Lacave, F. Sueur, On the motion of a small body immersed in a two

dimensional incompressible perfect fluid, Bull. Soc. Math. France 142 (2014), no.

2, 1–48.

[21] O. Glass, A. Munnier, F. Sueur, Dynamics of a point vortex as limits of a shrinking

solid in an irrotational fluid, preprint 2014, arXiv:1402.5387.

35

[22] O. Glass, L. Rosier, On the control of the motion of a boat, Math. Mod. Meth.

Appl. Sci. 23 (2013), no. 4, 617–670.

[23] O. Glass, F. Sueur, The movement of a solid in an incompressible perfect fluid as

a geodesic flow, Proc. Amer. Math. Soc. 140 (2012), 2155–2168.

[24] O. Glass, F. Sueur, Uniqueness results for weak solutions of two-dimensional fluid-

solid systems, Archive for Rational Mechanics and Analysis. Volume 218 (2015),

Issue 2, 907–944.

[25] O. Glass, F. Sueur, T. Takahashi, Smoothness of the motion of a rigid body im-

mersed in an incompressible perfect fluid, Ann. Sci. de l’Ecole Normale Superieure

Volume 45, fascicule 1 (2012), 1–51.

[26] K. A. Grasse, Nonlinear perturbations of control-semilinear control systems, SIAM

J. Control Optim. 20 (1982), No. 3, 311–327.

[27] K. A. Grasse, Perturbations of nonlinear controllable systems, SIAM J. Control

Optim. 19 (1981), No. 2, 203–220.

[28] A. Henrot, M. Pierre, Variation et optimisation de formes, Une analyse

geometrique, Springer 2005 (in French).

[29] J.-G. Houot, J. San Martin, M. Tucsnak, Existence and uniqueness of solutions

for the equations modelling the motion of rigid bodies in a perfect fluid. J. Funct.

Anal., 259(11):2856–2885, 2010.

[30] O. Imanuvilov, T. Takahashi, Exact controllability of a fluid-rigid body system. J.

Math. Pures Appl. 87, Issue 4 (2007), 408-437.

[31] R. Lecaros, L. Rosier, Control of underwater vehicles in inviscid fluids. I: Irrota-

tional flows, ESAIM Control Optim. Calc. Var. 20 (2014), no. 3, 662-703.

[32] J. Loheac, A. Munnier. Controllability of 3D low Reynolds number swimmers.

ESAIM: Control, Optimisation and Calculus of Variations, 20 (2014) no. 1, 236-

268.

[33] J. E. Marsden, T. Ratiu, Introduction to mechanics and symmetry: a basic exposi-

tion of classical mechanical systems (Vol. 17). Springer Science & Business Media.

2013.

[34] A. Munnier, Locomotion of Deformable Bodies in an Ideal Fluid: Newtonian versus

Lagrangian Formalisms. J. Nonlinear Sci (2009) 19: 665-715.

36

[35] J. Ortega, L. Rosier, T. Takahashi, On the motion of a rigid body immersed in

a bidimensional incompressible perfect fluid. Ann. Inst. H. Poincare Anal. Non

Lineaire, 24 (2007), no. 1, 139–165.

[36] J. H. Ortega, L. Rosier, T. Takahashi. Classical solutions for the equations mod-

elling the motion of a ball in a bidimensional incompressible perfect fluid. M2AN

Math. Model. Numer. Anal., 39 (2005), no. 1, 79–108.

[37] C. Rosier, L. Rosier. Smooth solutions for the motion of a ball in an incompressible

perfect fluid. Journal of Functional Analysis, 256 (2009), no. 5, 1618–1641.

[38] V. I. Yudovich, The flow of a perfect, incompressible liquid through a given region.

Dokl. Akad. Nauk SSSR 146 (1962), 561–564 (Russian). English translation in

Soviet Physics Dokl. 7 (1962), 789–791.

37


Recommended